• No se han encontrado resultados

Assessment of Fe2O3/SiO2 catalyst for the continuous treatment of phenol aqueous solutions in a fixed bed reactor

N/A
N/A
Protected

Academic year: 2020

Share "Assessment of Fe2O3/SiO2 catalyst for the continuous treatment of phenol aqueous solutions in a fixed bed reactor"

Copied!
17
0
0

Texto completo

(1)

ASSESSMENT OF Fe

2

O

3

/SiO

2

CATALYSTS FOR THE

CONTINUOUS TREATMENT OF PHENOL AQUEOUS

SOLUTIONS IN A FIXED BED REACTOR

J.A. Botas*, J.A. Melero, F. Martínez, M.I. Pariente

Department of Environmental and Chemical Technology, ESCET, Rey Juan Carlos University,

c/Tulipán s/n, 28933, Móstoles, Madrid, Spain

Published on:

Catalysis Today 149 (2010) 334–340

doi:10.1016/j.cattod.2009.06.014

Keywords: iron catalyst; mesostructured silica; phenol; fixed bed reactor; CWHPO

(2)

Abstract

Different iron-containing catalysts have been tested for the oxidation of phenol aqueous

solutions in a catalytic fixed bed reactor in presence of hydrogen peroxide. All the catalysts

consist of iron oxide, mainly crystalline hematite particles, over different silica supports

(mesostructured SBA-15 silica and non-ordered mesoporous silica). The immobilization of iron

species over different silica supports was addressed by direct incorporation of metal during the

synthesis or post-synthesis impregnation. The synthesis conditions were tuned up to yield

agglomerated catalysts with iron loadings between 10 and 15 wt. %. The influence of the

preparation method and the type of silica support was evaluated in a catalytic fixed bed reactor

for the continuous oxidation of phenol in terms of its activity (phenol and total organic carbon

degradation) as well as its stability (catalyst deactivation by iron leaching). Those catalyst

prepared by direct synthesis either in presence (Fe2O3/SBA-15(DS)) or absence

(Fe2O3/SiO2(DS)) of template molecules achieved high catalytic performances (TOC reduction

of 65 and 52 %, respectively) with remarkable low iron leaching in comparison with the

impregnated iron catalysts. Catalytic results demonstrated that the synthesis method plays a

crucial role in the dispersion and stability of active species and hence resulting in superior

(3)

1. Introduction

The degradation of aromatic pollutants in wastewater streams generated by industrial

processes has emerged as an important concern due to the restrictive environmental laws and

regulations. Phenol and phenol-like compounds show high toxicity for microorganisms, high

chemical oxygen demand and poor biodegradability. Advanced Oxidation Processes (AOPs) are

postulated as interesting alternatives for the destruction of organic pollutants in industrial

wastewaters. These processes involve the generation of non-selective and highly reactive

hydroxyl radical (●OH), which is one of the most powerful oxidation agents [1], higher than

others chemical oxidants commonly used in wastewater treatments.

The use of hydrogen peroxide as oxidant is an option commercially available for the

treatment of industrial wastewaters [2,3]. Nevertheless, the hydrogen peroxide by itself is not

effective at high concentrations of refractory compounds due to the reaction rate is very low,

even for high oxidant concentrations [4]. In order to promote the generation of hydroxyl radicals

and reduce secondary reactions, the use of catalytic processes, such as Fenton systems, could be

interesting alternatives. The reaction of hydrogen peroxide with ferrous salts (Fenton’s reagent)

or other low-valence transition metals (Fenton-like reactions) at room temperature are well

known sources of hydroxyl radicals [5]. In some cases, temperature has been increased up to

80-120 ºC leading to the so-called Catalytic Wet Hydrogen Peroxide Oxidation (CWHPO)

processes. The homogeneous oxidation of organic compounds by means of Fenton-like

reactions has been studied in numerous works during last years and it is commercially used to

treat industrial wastewaters [6]. However, these homogeneous systems present two main

drawbacks. Firstly, the hydroxyl radical production by Fenton-like reactions is strongly

dependent on limited pH range (2.5-3.5) and secondly, a final separation of soluble iron species

from the treated wastewater is needed.

These drawbacks have promoted the development of Fenton and CWHPO processes

based on heterogeneous catalytic systems. The catalyst design is usually a key factor in order to

increase its surface area, minimize the metal sintering, improve its chemical stability and govern

(4)

incorporation of active iron species over different supports, such as silica, zeolites [7-8],

hexagonal mesostructured materials (MCM-41, HMS 9 and SBA-15) [9-13], as well as pillared

clays [14] for its application in CWHPO. An interesting overview of solid catalysts is reported

by Perathoner and Centi [15], in which Al-Fe-clays are shown as one group of the most

promising catalysts. More recently, the same authors have stated that copper pillared clays are

highly effective for WHPO of wastewater streams from olive oil milling production with low

loss of copper by leaching [16]. Melero et al. [11] have also reported a remarkable catalytic

performance of a novel Fe2O3/SBA-15 composite catalyst as compared to homogeneous systems

as well as a low dependence on the pH for the treatment of phenolic aqueous solutions.

These heterogeneous processes have been carried out in batch operations and phenol is

frequently used as model pollutant. The use of solid catalysts for CWHPO indicates a good

perspective of this technology for its application in continuous processes, but it still requires

large engineering efforts in modelling and optimal designing of catalyst and industrial reactors.

Moreover, their application in continuous systems is normally questioned by the catalyst

fouling, in particular for the treatment of wastewater with a complex composition. On the other

hand, for its application in fixed bed reactors, catalysts must be shaped into bodies such as

granules, spheres and extrudates, in order to decrease the pressure drop and to achieve a

convenient mechanical strength. However, there is almost no information in the open literature

about its agglomeration process, although this is an essential step regarding its commercial

application. Therefore, few works are described in literature dealing with the treatment of

organic pollutants by means of a fixed bed reactor using hydrogen peroxide as oxidant. In this

sense, a major experience of continuous processes can be found in catalytic wet air oxidation. In

this context, the aim of this contribution has been the study of the influence of several supports

and iron incorporation methods for the degradation of phenol aqueous solutions using a

(5)

2. Experimental

2.1. Preparation of powder and extruded Fe/SiO2 materials

Several powder silica-supported iron materials were prepared following two different

methods: direct incorporation of the metal during the synthesis and post-synthesis incipient

wetness impregnation. In the first strategy, iron was incorporated by direct co-condensation of

the silica (TEOS) and iron (FeCl3·6H2O) sources during the synthesis of: i) mesostructured

SBA-15 materials with hexagonal mesoscopic order (Fe2O3/SBA-15(DS)) following the method

described elsewhere [17] and ii) non-ordered silica support synthesized by means of the sol-gel

method, based on acid-catalysed hydrolysis and basic condensation with ammonium hydroxide

(Fe2O3/SiO2(DS)). The second method based on the incipient wetness impregnation with

Fe(NO3)3·9H2O aqueous solution was carried out over: i) silica SBA-15 material (Fe2O3

/SBA-15(I)), SiO2 sol-gel (Fe2O3/SiO2(I)) and commercial silica from PQ Corporation (FeI/SiO2

-PQ(I)). With the purpose of comparison, a physical mixture of hematite iron oxide (Fe2O3;

Aldrich) with the mesostructured SBA-15 silica (Fe2O3+SBA-15(PM)) was also prepared. Prior

to the agglomeration of the powder catalysts, all the materials were calcined at 550 ºC for 5

hours.

Fe/SiO2 extrudates were prepared by blending of the powder catalyst with sodium

bentonite (75:25 wt %) and synthetic methylcellulose polymer (10 wt. % of the previous

mixture) which act as binders in the extrusion process [18]. All the components were kneaded

with deionised water under high shear conditions until obtaining a homogeneous paste, which

was conditioned in a damp atmosphere. Thereafter, the paste was pushed through a ram extruder

of 4.8 mm circular die and the resultant rod-shaped materials were dried for 3 days in a climatic

chamber under controlled temperature and relative humidity (RH) from 20 ºC and 70 % RH to

40 °C and 10 % RH. Finally, the extrudates were calcined in air until 650 °C (2 hours) using a

slow ramp of temperature, in order to achieve a gradual removal of water and the organic

content, strengthening the mechanical consistence of the material by sintering of the inorganic

binder. The final pellets for the catalytic reactions were crushed and sieved until particle sizes

(6)

2.2. Catalysts characterization

Physicochemical and textural properties as well as the presence of crystalline phases of

the powder materials and extruded catalysts were analyzed by means of different conventional

techniques. X-ray powder diffraction (XRD) data were acquired on a Philips X-Pert

diffractometer using Cu Kα radiation. The data were collected ranging 2θ from 0.5 to 50º with a

resolution of 0.02º. Nitrogen adsorption and desorption isotherms at 77 K were measured using

a Micromeritics Tristar 3000 System. Transmission electron microscopy (TEM) images were

taken on a Philips Tecnai-20 electron microscope operating at 200 kV. SEM images were

obtained in a XL30 ESEM FEI (PHILIPS) electron microscope. The iron content of the

synthesised materials was measured by means of Atomic Emission Spectroscopy with Induced

Coupled Plasma (ICP-AES) analysis collected in a Varian Vista AX system.

2.3. Catalytic tests in an up-flow fixed bed reactor

Fig. 1 depicts the experimental set-up used for the degradation of phenol aqueous

solutions by means of Catalytic Wet Hydrogen Peroxide Oxidation (CWHPO). The tubular

reactor was made of glass with inner diameter of 1.2 cm and 15 cm of length. The catalyst

particles were packed between glassy beads to enable a better distribution of the inlet solution

inside the catalyst bed and achieve the desired temperature. The operation conditions for the

evaluation of the activity and stability of the synthesized materials were set according to

preliminary studies with this experimental set-up [19]: i) reactor temperature of 80 ºC, ii) 1

mL/min of feed flow rate, iii) 1 g/L of phenol concentration and 5.1 g/L of H2O2 (stoichiometric

amount for the complete phenol mineralization, according to the reaction (1)) in the inlet

solution, and iv) 2.9 g of catalyst with a particle size between 1.6 and 2 mm.

(7)

Pump

FBR

Thermal bath

Cooler

Sample

Treated effluent TR

FC

TI TC

QA Feed

110 ºC

SET MENU

Pump

FBR

Thermal bath

Cooler

Sample

Treated effluent TR

FC

TI TC

QA Feed

110 ºC

SET MENU

Fig. 1.Experimental set-up of the up-flow fixed bed reactor for the catalytic wet hydrogen peroxide oxidation of phenol aqueous solutions.

Samples were hourly withdrawn from the treated effluent to determine Total Organic

Carbon (TOC), phenol and hydrogen peroxide concentrations, as well as iron leached out from

the catalyst to the treated effluent. TOC content of the solutions was analysed using a

combustion/non dispersive infrared gas analyser model TOC-V Shimadzu. Phenol conversion

was determined by means of HPLC chromatograph (Varian Prostar) equipped with a Waters

Spherisorb column and an UV detector adjusted at 215 nm, employing ultrapure water acidified

with H3PO4 up to pH 2.7 as mobile phase. Hydrogen peroxide conversion was determined by

iodometric titration. Iron content in the treated solution was measured by inductively coupled

plasma-atomic emission spectroscopy (ICP-AES) analysis collected in a Varian VISTA AX

(8)

3. Results and discussion

3.1. Catalysts characterization

Table 1 shows the physicochemical properties of the synthesized powder and extruded

iron catalysts, as well as the silica materials used as supports for impregnation. The materials

have been classified in different groups depending on the silica support: mesostructured

SBA-15, mesoporous silica obtained via sol-gel and commercial extruded mesoporous silica acquired

from PQ Corporation. As can be seen, the first group corresponds to hexagonal mesostructured

SBA-15 materials, where iron was incorporated by direct synthesis (Fe2O3/SBA-15(DS)) and

incipient wetness impregnation (Fe2O3/SBA-15(I)). Additionally, physical mixture of hematite

iron oxide and the SBA-15 silica support (Fe2O3+SBA-15(PM)) was also prepared. The second

group includes mixed Fe2O3-SiO2 materials synthesized by simultaneous condensation of iron

and silica sources via the sol-gel method (Fe2O3/SiO2(DS)) or impregnation of mesoporous

silica sol-gel (Fe2O3/SiO2(I)). Finally, the third group is constituted by impregnated commercial

silica support (Fe2O3/SiO2-PQ(I)).

Table 1. Characterization data of iron-containing catalysts supported over silica materials.

Material

(Synthesis method) Support

Powder Extruded

SBET

(m2/g) (cmVp 3/g) Fe (wt.%) (mSBET 2/g) (cmVp 3/g) Fe (wt.%)

SBA-15

SBA-15

790 1.21 - - -

-Fe2O3/SBA-15(I) 549 0.78 15.8 333 0.53 12.4

Fe2O3/SBA-15(DS) 475 0.67 18.0 264 0.48 14.0

Fe2O3+SBA-15(PM) - - - 277 0.49 13.8

SiO2

SiO2

sol-gel

296 1.79 -

-Fe2O3/SiO2(I) 207 0.85 15.9 180 0.62 12.0

Fe2O3/SiO2(DS) 270 0.69 14.8 225 0.55 11.1

SiO2-PQ

SiO2 PQ

- - - 346 1.16

-Fe2O3/SiO2–PQ(I) - - - 252 0.87 15.4

Characterization data of powder materials evidence that catalysts based on SBA-15 silica

supports have higher specific surface areas than those supported over the silica synthesized via

(9)

nm) with hexagonal arrangement, which provides remarkable surface areas and pore volumes

(650-800 m2/g and 1-1.2 cm3/g, respectively). In contrast, silica sol-gel materials depict broader

pore size distributions and higher average pore diameter than SBA-15 based materials, which

lead to higher ratio between pore volume and specific surface area. As result of the iron

incorporation process over both silica supports, a significant reduction of surface area and pore

volume were evidenced for all the catalysts, although this fact was particularly more

accentuated for the incorporation of iron by direct synthesis over SBA-15 materials

(Fe2O3/SBA-15(DS)) [20]. Nevertheless, the powder catalysts synthesized over SBA-15

supports show higher specific surface areas regardless of the synthesis method (475 and 550

m2/g, for Fe

2O3/SBA-15(DS) and Fe2O3/SBA-15(I), respectively) than the counterparts

supported over non-ordered mesoporous silica (270 and 207 m2/g for Fe

2O3/SiO2(DS) and

Fe2O3/SiO2(I), respectively).

XRD patterns illustrated in Fig. 2 evidence typical diffractions of the hexagonal

arrangement of mesoporous materials for powder iron containing SBA-15 catalysts in the low

angle range, whereas iron catalyst prepared by gel method or impregnation over silica

sol-gel does not show any sign of mesoscopic order. At higher angles, clear diffraction peaks at 33,

36 and 42 º are observed, which are characteristic of crystalline hematite iron oxide particles,

except for Fe2O3/SBA-15(I) catalyst. This fact could be related to the small size of the

crystalline iron oxide particles which makes less visible to the X ray diffraction. The

comparison between XRD pattern of powder and extruded materials indicates that the

characteristic diffractions of hematite iron oxide particles are not affected by the agglomeration

process. Additionally, the hexagonal mesoscopic order of SBA-15 materials was also

maintained.

As can be seen in Table 1, the percentage of iron incorporated over the powder catalysts

ranged from 15 to 18 wt. %, whereas for agglomerated materials decreased up to 11-15 wt. %.

This reduction is consequence of the dilution effect related to the inorganic bentonite clay

incorporated into the extruded material during the agglomeration process. Moreover, the

(10)

the powder counterparts. This phenomenon is also attributed to the amount of bentonite clay

incorporated in the extruded material, which presents a low surface area (19 m2/g), as well as

the sintering process that take place during the final calcination step. Interestingly, the effect of

the sintering process seems to be more important in SBA-15-based iron catalysts, which suffer

40-45 % of surface area reduction, than iron catalyst supported over non-ordered silica sol-gel

materials (with reductions in the range 13-16 %). Nevertheless, iron containing catalysts based

on mesostructured SBA-15 silica matrices still exhibit higher specific surface areas. Pore size

distribution profiles shown in Fig. 3 for agglomerated catalysts confirm the characteristic pore

size distribution of mesostructured SBA-15-type materials with a narrow pore size distribution

centred at 8-9 nm, as compared to those shown by non-ordered mesoporous materials prepared

by the sol-gel method with a broader pore size distribution. Unlike the physical mixture of

Fe2O3+SBA-15(PM) which have shown the highest pore diameter (8.8 nm), the iron

incorporation over SBA-15 silica support by direct synthesis or impregnation have yielded a

slight reduction of the pore diameter up to ca. 7.7 and 8.2 nm, respectively. Regarding to iron

catalysts supported over non-ordered mesoporous silica, the sample Fe2O3/SiO2(DS) exhibits a

broader pore size distribution centred at ca. 15 nm as compared to the other two catalysts of this

group.

TEM images of extruded iron catalysts (Fig. 4) confirm the hexagonal arrangement of

SBA-15-based materials. In both groups of iron containing catalysts, mesostructured SBA-15

and non-ordered mesoporous silica materials, the iron particle size is strongly dependent on the

synthesis method and type of silica support. Interestingly, Fe2O3/SiO2(DS) catalyst, prepared by

the sol-gel method, evidence a remarkable dispersion of iron particles over the SiO2 support.

Unlike iron impregnated catalysts over non-ordered mesoporous silica sol-gel (Fe2O3/SiO2(I))

and silica-PQ (Fe2O3/SiO2-PQ(I)), impregnated Fe2O3/SBA-15(I) catalyst shows a better

dispersion of smaller iron particles over the silica framework. These data seem to be in

agreement with the small iron oxide particle sizes assumed from the X ray diffraction patterns.

The presence of agglomerating bentonite layers can be also observed together with the silica

(11)

(b)

Fe2O3+SBA-15(PM)

Fe2O3/SBA-15(I)

Fe2O3/SBA-15(DS)

25 30 35 40 45

2θ

1 2

Fe2O3/SiO2-PQ(I)

Fe2O3/SiO2(I)

Fe2O3/SiO2(DS)

2θ (º)

1 2

Fe2O3/SiO2(I)

Fe2O3/SiO2(DS)

SiO2

2θ (º)

25 30 35 40 45

2θ (a)

Fe2O3/SBA-15(I)

Fe2O3/SBA-15(DS)

SBA-15 In te ns ity (a .u ) (b)

Fe2O3+SBA-15(PM)

Fe2O3/SBA-15(I)

Fe2O3/SBA-15(DS)

25 30 35 40 45

2θ

1 2

Fe2O3/SiO2-PQ(I)

Fe2O3/SiO2(I)

Fe2O3/SiO2(DS)

2θ (º)

1 2

Fe2O3/SiO2(I)

Fe2O3/SiO2(DS)

SiO2

2θ (º)

25 30 35 40 45

2θ (a)

Fe2O3/SBA-15(I)

Fe2O3/SBA-15(DS)

SBA-15 In te ns ity (a .u )

Fig. 2. XRD patterns of (a) powder and (b) extruded iron-containing materials over mesostructured

SBA-15 and non ordered mesoporous silica supports.

1 10 100

0,00 0,05 0,10 0,15 0,20 0,25 0,30 0,35 0,40

Fe2O3/SiO2(DS)

Fe2O3/SiO2(I)

Fe2O3/SiO2-PQ(I)

(b) d V /d lo g( D ) ( cc )/ g-n m )

Pore diameter (nm)

1 10 100

0,00 0,05 0,10 0,15 0,20 0,25 0,30 0,35 0,40

Fe2O3/SBA-15(DS)

Fe2O3/SBA-15(I)

Fe2O3+SBA-15(PM)

(a) d V /d lo g( D ) (c c) /g -n m )

Pore diameter (nm)

1 10 100

0,00 0,05 0,10 0,15 0,20 0,25 0,30 0,35 0,40

Fe2O3/SiO2(DS)

Fe2O3/SiO2(I)

Fe2O3/SiO2-PQ(I)

(b) d V /d lo g( D ) ( cc )/ g-n m )

Pore diameter (nm)

1 10 100

0,00 0,05 0,10 0,15 0,20 0,25 0,30 0,35 0,40

Fe2O3/SBA-15(DS)

Fe2O3/SBA-15(I)

Fe2O3+SBA-15(PM)

(a) d V /d lo g( D ) (c c) /g -n m )

Pore diameter (nm)

Fig. 3. Pore size distribution of the extruded materials: (a) supported over SBA-15; (b) supported over non-ordered mesoporous SiO2.

Fe2O3

SiO2

1 µm

SBA-15

Bentonite Fe2O3

200 nm

Fe2O3+SBA-15(PM)

500 nm

Fe2O3

SBA-15 Fe2O3/SBA-15(DS)

200 nm

Fe2O3-SiO2

500 nm

SBA-15

Fe2O3

1 µm

SiO2

Fe2O3

Fe2O3/SiO2-PQ(I)

Fe2O3/SBA-15(I)

Fe2O3/SiO2(DS) Fe2O3/SiO2(I)

Fe2O3

SiO2

1 µm

Fe2O3

SiO2

1 µm

SBA-15

Bentonite Fe2O3

200 nm

Fe2O3+SBA-15(PM)

500 nm

Fe2O3

SBA-15

500 nm 500 nm

Fe2O3

SBA-15 Fe2O3/SBA-15(DS)

200 nm

Fe2O3-SiO2

500 nm

SBA-15

Fe2O3

1 µm

SiO2

1 µm

SiO2

Fe2O3

Fe2O3/SiO2-PQ(I)

Fe2O3/SBA-15(I)

Fe2O3/SiO2(DS) Fe2O3/SiO2(I)

(12)

3.2. Catalytic treatment of phenolic aqueous solutions in a fixed bed reactor

The catalytic performance of iron-containing materials for the wet peroxide oxidation of

phenol aqueous solutions was evaluated in terms of TOC reduction, phenol degradation and

hydrogen peroxide conversion. Likewise, the stability of the catalysts along the treatment was

also studied analyzing the iron concentration leached out from the catalyst within the treated

outlet effluent. Table 2 summarizes the results of the parameters monitored in the outlet stream

at steady-state after eight hours of time on stream for all the synthesized catalysts.

Table 2. Activity and stability of the different iron-containing catalyst under steady-state conditions.

Catalyst XTOC

(%) X(%)Phenol X(%)H2O2 η∗ [Fe](mg/L)leached pH

Fe2O3/SBA-15(I) 61 100 100 0.52 123 4.0

Fe2O3/SBA-15(DS) 64 100 100 0.56 14 4.3

Fe2O3+SBA-15(PM) 62 100 100 0.53 5 5.0

Fe2O3/SiO2(I) 64 100 100 0.54 20 4.2

Fe2O3/SiO2(DS) 52 100 100 0.44 11 4.0

Fe2O3/SiO2-PQ(I) 21 93 100 0.18 120 2.3

η = Theoretical H

2O2 consumption/ Real H2O2 consumption

It is noteworthy that phenol is completely degraded for all the catalysts regardless of the

silica support and synthesis method. Results of TOC evidence a high catalytic performance of

SBA-15-based iron catalysts with conversions above 60 %. In the case of sol-gel materials,

Fe2O3/SiO2(I) keep a similar TOC conversion (64 %), whereas Fe2O3/SiO2(DS) undergoes a

slight reduction up to 52 %. In contrast, impregnated commercial silica PQ (Fe2O3/SiO2-PQ(I))

shows a striking decrease of TOC degradation until ca. 25 %. The hydrogen peroxide is totally

consumed at the operation conditions under study for all the catalysts and with similar

efficiency (η≈0.55) except for the catalysts with lower TOC conversion previously mentioned

(Fe2O3/SiO2(DS) and Fe2O3/SiO2-PQ(I)). Values of oxidant efficiency around 0.5-0.6 indicate an

acceptable use of the oxidant. It must be pointed out that oxygenated by-products (mainly

carboxylic acids) formed at high TOC degradation rates are much more refractory than phenolic

(13)

Nevertheless, it is particularly remarkable the significant efficiency decrease for Fe2O3/SiO2

-PQ(I) (η = 0.18). The lowest oxidant efficiency of Fe2O3/SiO2-PQ(I) could be related to the

nature of the commercial silica or presence of metal traces that enable the decomposition of the

oxidant to other less active radical species, different of hydroxyl radicals.

Regarding iron leaching, it is important to note that catalysts prepared by impregnation

(Fe2O3/SiO2(I), Fe2O3/SiO2-PQ(I) and Fe2O3/SBA-15(I)) display higher loss of iron in the outlet

stream than those prepared by direct synthesis (Fe2O3/SiO2(DS) and Fe2O3/SBA-15(DS)). This

fact indicates weaker interactions between the iron species and the silica support for

iron-containing catalysts prepared by impregnation, which is not only contributed to faster catalyst

deactivation but also to additional metal pollution in the treated effluent, as it is the case of

Fe2O3/SBA-15(I) and Fe2O3/SiO2-PQ(I) catalysts with iron concentrations in the outlet stream

above 100 mg/L. The best stability of Fe2O3/SiO2(I), as compared to the other two impregnated

catalyst over mesostructured SBA-15 and mesoporous silica-PQ, indicates a stronger interaction

between iron particles and silica support. This can be related to the particular physicochemical

properties of the silica matrix for supporting small iron clusters. In this sense, the high density

of silanol groups that typically provides silica prepared via sol-gel, could be responsible of

stronger interlinking of the impregnated iron species over the silica framework.

Looking at the pH values, the decrease of the initial phenol aqueous solution pH (ca. 5.5)

is observed. This decrease is normally attributed to the formation of carboxylic acids as final

refractory by-products of the partial oxidation of phenol as well as protons by Fenton related

reactions [22]. Some studies have reported that metal iron leaching of Fenton-like catalysts can

be influenced by the solution pH in systems operating in batch conditions [12]. In this study for

the continuous treatment of phenol aqueous solution with different catalysts this type of

relationship was not evidenced.

The feasibility and durability of the iron catalysts for the continuous treatment of a phenol

aqueous solution can be also demonstrated by the results of TOC conversion and iron leaching

along the time on stream (Figure 5 and 6, respectively). TOC conversion profiles revealed times

(14)

mixture of Fe2O3+SBA-15(PM), which needs almost 7 hours. Although this enlargement of the

stabilizing time has not deeply studied, this phenomenon might be related to particular

diffusional constraints associated with the accessibility of the supported iron particles depending

on the silica support and the type of synthesis method. Nevertheless, it must be remarked the

stability of the TOC degradation rate, once the steady state is reached, up to 8 hours on stream

for all the catalyst, excepting Fe2O3+SBA-15(PM).

0 1 2 3 4 5 6 7 8

0 10 20 30 40 50 60 70 80 90 100 (a)

XTO

C

(

%

)

Time on stream (hours) 0 1 2 3 4 5 6 7 8

0 10 20 30 40 50 60 70 80 90 100 (b)

XTO

C

(

%

)

Time on stream (hours)

0 1 2 3 4 5 6 7 8

0 10 20 30 40 50 60 70 80 90 100 (a)

XTO

C

(

%

)

Time on stream (hours)

0 1 2 3 4 5 6 7 8

0 10 20 30 40 50 60 70 80 90 100 (a)

XTO

C

(

%

)

Time on stream (hours) 0 1 2 3 4 5 6 7 8

0 10 20 30 40 50 60 70 80 90 100 (b)

XTO

C

(

%

)

Time on stream (hours)

0 1 2 3 4 5 6 7 8

0 10 20 30 40 50 60 70 80 90 100 (b)

XTO

C

(

%

)

Time on stream (hours)

Fig. 5. TOC conversion profiles for the different iron-containing catalysts prepared by means of direct synthesis and incipient wetness impregnation: (a) -□- Fe2O3/SBA-15(I), -◄- Fe2O3/SBA-15(DS),

-○- Fe2O3+SBA-15(PM); (b) -▼- Fe2O3/SiO2(I), -◊- Fe2O3/SiO2(DS), -■- Fe2O3/SiO2-PQ(I).

According to the leached iron profiles (Figure 6), catalysts prepared by direct synthesis

Fe2O3/SBA-15(DS) and Fe2O3/SiO2(DS) as well as impregnated Fe2O3/SiO2(I) exhibited a good

behaviour with relatively low values, between 10 and 20 mg/L, and slight variation of the iron

concentration detected along the time on stream. Profiles of Fe2O3/SBA-15(I) and Fe2O3/SiO2

-PQ(I) catalysts, evidence a low durability for long operating times. Attending to the legal

discharge limit for iron concentration (10 mg/L), one of the main aims to apply successfully

heterogeneous catalytic systems is the reduction of iron leaching to lower values, which

promote higher durability of the catalyst. In this sense, the physico-chemical properties of the

materials prepared by direct synthesis, independently of the used support, allow improving the

catalytic stability, being consequently more interesting for this kind of applications.

Summarizing, catalysts prepared by direct synthesis over mesostructured SBA-15 and

(15)

of TOC reduction, and stability, measured as iron leached out to the treated effluent. These

materials provide activities above 50 % in terms of TOC conversion as well as reduced iron

leaching below 15 mg/L. Although mesostructured Fe2O3/SBA-15(DS) catalyst presents higher

values of TOC degradation (64 %) and hydrogen peroxide efficiency, the non-ordered

mesoporous Fe2O3/SiO2(DS) prepared via sol-gel is also considered as a potential candidate for

its application in industrial catalytic wet peroxide oxidation processes. The decreased catalytic

performance of Fe2O3/SiO2(DS) as compared to its counterpart over mesostructured SBA-15

support could be related to its either lower iron content (11.1 vs 14 %) or lower specific surface

area (225 vs 264 m2/g).

0 1 2 3 4 5 6 7 8

0 20 40 60 80 100 120 140 160 180 (a) [F e]le ac h ed ( m g/ L )

Time on stream (hours)

0 1 2 3 4 5 6 7 8

0 20 40 60 80 100 120 140 160 180 (b) [F e]le ac h ed ( m g /L )

Time on stream (hours)

0 1 2 3 4 5 6 7 8

0 20 40 60 80 100 120 140 160 180 (a) [F e]le ac h ed ( m g/ L )

Time on stream (hours)

0 1 2 3 4 5 6 7 8

0 20 40 60 80 100 120 140 160 180 (b) [F e]le ac h ed ( m g /L )

Time on stream (hours)

Fig. 6. Leached iron concentration for the different iron/silica catalysts prepared by means of direct synthesis and incipient wetness impregnation: (a) -□- Fe2O3/SBA-15(I), -◄- Fe2O3/SBA-15(DS),

-○- Fe2O3+SBA-15(PM); (b) -▼- Fe2O3/SiO2(I), -◊- Fe2O3/SiO2(DS), -■- Fe2O3/SiO2-PQ(I).

4. Conclusions

Different iron-containing catalysts based on hematite crystalline iron oxides supported

over mesostructured SBA-15 and non-ordered mesoporous silica were tested in an up flow fixed

bed reactor for the continuous catalytic wet peroxide oxidation of phenol aqueous solution.

SBA-15 silica support provides iron catalysts with narrower pore size distributions and higher

specific surface areas than non-ordered mesoporous silica. The agglomeration process do not

produce significant changes in the physicochemical properties of the precursor powder

catalysts, although in the case of mesostructured SBA-15 silica supports, a more accentuated

(16)

The activity of the iron containing catalyst in terms of TOC conversion and oxidant

efficiency was mainly affected by the type of silica support, whereas the iron resistance to be

leached out was principally dependent on the synthesis method. The iron incorporation in the

silica matrix by means of direct synthesis (co-condensation of iron and silica sources) has

provided higher stability of the supported iron species. In contrast, those catalysts prepared by

incipient wetness impregnation proved higher losses of iron in the treated aqueous solution (up

to 100 mg/L). Mesostructured Fe2O3/SBA-15 is shown as the best catalyst with a high catalytic

performance and reduced iron leaching for the treatment of phenolic aqueous solutions.

Non-ordered mesoporous Fe2O3/SiO2(DS) catalyst prepared via sol-gel in absence of organic

template was also considered an interesting candidate for its application in wet peroxide

oxidation processes. Both catalysts achieved a total phenol removal with TOC reductions above

50 % and relative low iron leaching (below 15 mg/L) operating for eight hours on stream.

Acknowledgements

The authors thank “Ministerio de Ciencia y Tecnología” for the financial support through

the project Consolider-Ingenio 2010 and “Comunidad de Madrid” through the project

P-AMB-000395-0505.

References

[1] W.H. Glaze, J.W. Kang, D.H. Chapin. Ozone: Sci. Eng. 9 (1987) 335.

[2] C.A. Tolman, W. Tumas, S.Y.L. Lee, D. Campos. D.H.R. BARTON, Plenum Press, New

York, 1993, p. 57.

[3] L. Plant, M. Jeff. Chem. Eng. 16 (1994) 16.

[4] E. Neyens, J. Baeyens. J. Hazard. Mater. 98 (2003) 33.

[5] H.J. Fenton. J. Chem. Soc. 65 (1894) 899.

[6] C.H. Walling. Acc. Chem. Res. 8 (1975) 125.

[7] G. Ovejero, J.L. Sotelo, F. Martínez, J.A. Melero, L. Gordo. Ind. Eng. Chem. Res. 40 (2001)

(17)

[8] J.A. Melero, G. Calleja, F. Martínez, R. Molina, K. Lázár. Microporous Mesoporous Mater.

74 (2004) 11.

[9] N. Crowther, F. Larachi. Appl. Catal. B: Environ. 46 (2003) 293.

[10] G. Calleja, J.A. Melero, F. Martínez, R. Molina. Water Res. 39 (2005) 1741.

[11] J.A. Melero, G. Calleja, F. Martínez, R. Molina, M.I. Pariente. Chem. Eng. J. 131 (2007)

245.

[12] F. Martínez, M.I. Pariente, J.A. Melero, J.A. Botas. J. Adv. Oxid. Technol. 11(2008) 75.

[13] M.I. Pariente, F. Martínez, J.A. Melero, J.A. Botas, T. Velegraki, N.P. Xekoukoulotakis, D.

Mantzavinos. Appl. Catal. B: Environ. 85 (2008) 24.

[14] E. Guélou, J. Barrault, J. Fournier, J.M. Tatibouët. Appl. Catal. B: Environ. 44 (2003) 1.

[15] S. Perathoner, G. Centi. Topics in Catal. 33 (2005) 207.

[16] S. Caudo, G. Centi, Ch. Genovese, S. Perathoner. Appl. Catal. B: Environ. 70 (2007) 437.

[17] D. Zhao, J. Feng, Q. Huo, N. Melosh, G.H. Fredrickson, B.F. Chmelka, G.D. Stucky.

Science 279 (1998) 548.

[18] J.R. González-Velasco, M.A. Gutierréz-Ortiz, R. Ferret, A. Aranzabal, J.A. Botas. J. Mater.

Sci. 34 (1999) 1999.

[19] F. Martínez, J.A. Melero, J.A. Botas, M.I. Pariente, R. Molina. Ind. Eng. Chem. Res. 46

(2007) 4396.

[20] L. Vradman, M.V. Landau, D. Kantorovich, Y.A. Koltypin, Microporous Mesoporous

Mater. 79 (2005) 307.

[21] G. Centi, S. Perathoner, T. Torre, M.G. Verduna. Catal. Today. 55 (2000) 61.

[22] M. Noorjahan, V. Durga Kumari, M. Subrahmanyam, L. Panda. Appl. Catal. B: Environ.

Referencias

Documento similar

set-up with the portable Raman and the gas cell for measurements are presented in Fig. a) Scheme of the experimental set-up used for detection of CWA vapours. b) Digital

The catalytic wet peroxide oxidation (CWPO) of bisphenol A (BPA) with Fe catalysts supported on activated carbon from grape seeds (GS) has been studied.. Oxidized samples

104 In this work, the capability of dark heterogeneous Fenton oxidation, also known as catalytic 105 wet peroxide oxidation (CWPO), for the elimination of the neonicotinoid

Initially phenol oxidation process was evaluated with the use of a copper salt as homogeneous catalyst, from where it was obtained the highest catalytic capacity of Cu(II) ions

Short term CWAO experiments showed that liquid flow modulation has only a mild although negative influence on phenol conversion as the cycle mean conversions are always lower than

In particular, chelate or intramolecular cooperativity is responsible for the assembly of closed, cyclic structures in competition with open, linear oligomers, and it can

Fairly high mineralization percentages (75% and 70%, using the magnetic and the non- magnetic catalysts, respectively) were achieved at the end of the experiments (4 h), while

In the current work, the effect of several operating conditions, including initial pH (3.5–10.5), catalyst load (0–6 g/L), initial phenol concentration (0.1–5 g/L), hydrogen